Hostname: page-component-8448b6f56d-m8qmq Total loading time: 0 Render date: 2024-04-16T18:53:17.127Z Has data issue: false hasContentIssue false

100 Million Years of Reef Prosperity and Collapse: Ordovician to Devonian Interval

Published online by Cambridge University Press:  21 July 2017

Paul Copper*
Affiliation:
Loupicoubas, 46220 Prayssac, France
Get access

Abstract

From the beginning of the Late Ordovician (Sandbian: 460.9myr) through end Devonian (Famennian: 359.2myr), coral-stromatoporoid sponge reefs formed a remarkable, evolving ecosystem that dominated sediment production on tropical carbonate platforms in a calcitic ocean. This was a time of maximal and unparalleled reef development in the Phanerozoic, with reef tracts vastly exceeding in size and biodiversity of those in the Holocene (e.g., the Great Barrier Reef). Within this circum-equatorial niche, the calcitic tabulate and rugose corals, and the aragonitic (or high Mg calcite) stromatoporoid sponges, were the primary Middle Paleozoic reef frame builders. These were supplemented ecologically and skeletally by now extinct groups of calcitic bryozoans, crinoids, brachiopods, and red algae, alongside aragonitic green algae, and enigmatic CaCO3 precipitating and binding calcimicrobes. This 100 myr long Middle Paleozoic reef consortium thrived under SST averages of 30°+, to latitudes as high as 45°–55°, under high atmospheric CO2 conditions of 6000+ ppm, and sealevels 150–200 m higher than today. This reef ecosystem was disrupted by several relatively short duration south polar glacial episodes, centered around northern Gondwana, defining the O/S boundary Mass Extinction Events (MEEs). Nearly all coral and stromatoporoid families survived this MEE: there were losses at the genus level. Reef-building stopped nearly everywhere, and during the ‘recovery’ interval, solitary rugose corals initially prevailed, and stromatoporoids were small. Full global re-establishment of the reef ecosystem, and biodiversity, took another 3–4 million years (not until the late Aeronian, Early Silurian). This was followed by a remarkable reef expansion in the Middle Silurian (Wenlock), then by declines in the latest Silurian (Ludlow-Pridoli), and earliest Devonian (Lochkovian) possibly due to sealevel lowstands, tectonic plate re-assembly, and ocean current re-direction. Maximal Phanerozoic reef success was during the Emsian-Givetian, when some 15 barrier reef tracts more than 1100 km long flourished in tropical shallow seas. Reef-building coral diversity exceeded 200 genera, and the calcifying stromatoporoids evolved 60+ genera, especially in the ‘Old World’ faunal province (Euramerica, Cathaysia, northern Australia). Near the end of the Middle Devonian (mid- to late Givetian), the primary reef dwellers declined sharply in diversity, marked generally by sealevel lowstand, followed in the Frasnian (Late Devonian) by shrinking latitudes for carbonate platforms, and reduced reef accommodation space. Sharp cooling, with the arrival of a global Icehouse climate, and aragonitic oceans, led to the second largest Phanerozoic Mass extinction around the Frasnian/Famennian boundary, with reef builder and reef inhabitant losses exceeding those of the O/S MEE. The global absence of coral-sponge reefs persisted for nearly all of the 16 myr long Famennian, as total CaCO3 production fell some 60–90%, as aragonitic oceans took over. Only small and scattered Famennian coral-stromatoporoid patch reefs are known, with the last of these in the late Famennian (Strunian), punctuated by total disappearance of the whole keystone reef-building order. Famennian and Strunian corals belonged to Carboniferous families. During the Famennian, calcimicrobes, the first calcifying foraminiferans, and select ‘lithistid’ calcareous sponges dominated a highly stressed reef ecosystem, lacking barrier reef tracts. Biodiversity and reef construction were decoupled under global climatic stress during the succeeding icehouse Late Paleozoic.

Type
Research Article
Copyright
Copyright © 2011 by The Paleontological Society 

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Algeo, T.J., Berner, R.A., Maynard, J.B., and Scheckler, S.E. 1995. Late Devonian oceanic anoxic events and biotic crises: ‘rooted’ in the evolution of vascular plants? Geology Today, 5:6366.Google Scholar
Alroy, J. 2008. Dynamics of extinction and origination in the fossil record. Proceedings of the National Academy of Sciences, USA, 105:1153611542.CrossRefGoogle Scholar
Alroy, J. 2010a. The shifting balance of diversity among major marine animal groups. Science, 329:11911194.CrossRefGoogle ScholarPubMed
Alroy, J. 2010b. Geographical, environmental and intrinsic biotic controls on Phanerozoic marine diversification. Palaeontology, 53:12111235.CrossRefGoogle Scholar
Alvaro, J.J., Aretz, M., Boulvain, F., Munnecke, A., Vachard, D., and Vennin, E. (eds.). 2007. Palaeozoic reefs and bioaccumulations: climatic and evolutionary controls. Geological Society, London Special Publications, 275:1291.Google Scholar
Aretz, M., and Chevalier, E. 2007. After the collapse of stromatoporid-coral reefs - the Famennian and Waulsortian reefs of Belgium: much more than Waulsortian mounds. In Alvaro, J.J., Aretz, M., Boulvain, F., Munnecke, A., Vachard, D., and Vennin, E. (eds.), Palaeozoic reefs and bioaccumulations: climatic and evolutionary controls. Geological Society, London Special Publications, 275:163188.CrossRefGoogle Scholar
Averbuch, O., Tribovillard, N., Devleeschouwer, X., Riquier, L., Mistiaen, B., and Van Vliet-Lanoe, B. 2005. Mountain-building-enhanced continental weathering and organic carbon burial as major causes for climatic cooling at the Frasnian-Famennian boundary. Terra Nova, 17:2534.CrossRefGoogle Scholar
Bambach, R.K. 2006. Phanerozoic biodiversity mass extinctions. Annual Review of Earth and Planetary Sciences, 34:127155.CrossRefGoogle Scholar
Bancroft, B.B. 1928. On the unconformity at the base of the Ashgillian in the Bala District. Geological Magazine, 65:484493.CrossRefGoogle Scholar
Barnosky, A.D., and 11 co-authors. 2011. Has the Earth's sixth mass extinction already arrived? Nature, 471:5157.CrossRefGoogle ScholarPubMed
Belenitskaya, G.A., and Zadoroshnaya, N.M. (eds.). 1990. Rifogennye isulfatonosnye formatsii Fanerozoya SSSR [Phaneroic reefal and sulfatic formations of the USSR], Moskva, Nedra, 292 pp.Google Scholar
Berner, R.A. 2004. The Phanerozoic carbon cycle: CO2 and O2 . Yale University Press, New Haven, 440 pp.CrossRefGoogle Scholar
Berner, R.A., Vandenbrooks, J.M., and Ward, P.D. 2007. Oxygen and evolution. Science, 316:537538.CrossRefGoogle ScholarPubMed
Brenchley, P.J. 2004. End Ordovician glaciation, pp. 8183 In Webby, B.D., Paris, F., Droser, M.L., and Percival, I.G. (eds.), The great Ordovician biodiversification event, Columbia University Press, New York, 484 pp.CrossRefGoogle Scholar
Budd, A. 2000. Diversity and extinction in the Cenozoic history of Caribbean reefs. Coral Reefs, 19:2535.CrossRefGoogle Scholar
Buggisch, W., and Joachimski, M.M. 2006. Carbon isotope stratigraphy of the Devonian of Central and Southern Europe. Palaeogeography, Palaeoclimatology, Palaeoecology, 240:6888.CrossRefGoogle Scholar
Buggisch, W., Joachimski, M.M., Lehnert, O., Bergstrom, S.M., Repetski, J., and Webers, G.F. 2010. Did intense volcanism trigger the first Late Ordovician icehouse? Geology, 38:327330.CrossRefGoogle Scholar
Caputo, M. V. 1985. Late Devonian glaciation in South America. Palaeogeography, Palaeoclimatology, Palaeoecology, 51:291317.CrossRefGoogle Scholar
Caputo, M.V., Melo, J.H.G., Sstreel, M., and Isbel, J. 2008. Late Devonian and Early Carboniferous glacial records. Geological Society of America Special Papers, 441:161173.Google Scholar
Carrera, M.G., and Rigby, J.K. 2004. Sponges, pp. 102111 In Webby, B.D., Paris, F., Droser, M.L., and Percival, I.G. (eds.), The great Ordovician biodiversification event, Columbia University Press, New York, 484 pp.CrossRefGoogle Scholar
Cecile, M.P. 1988. Ordovician reefs and organic buildups. In Geldsetzer, H.H.J., James, N.P., and Tebbutt, G.E. (eds.), Reefs - Canada and adjacent areas. Memoirs Canadian Society of Petroleum Geologists, 13:171176.Google Scholar
Chen, X, Rong, J., Fan, J., Zhan, R., Mitchell, C.E., Harper, D.A.T., Melchin, M.J., Peng, P., Finney, S.C., and Wang, X. 2006. The global boundary stratotype section and point (GSSP) for the base of the Hirnantian Stage (the uppermost of the Ordovician System). Episodes, 29:183196.CrossRefGoogle Scholar
Cocks, L.R.M., Holland, C.H., Rickards, R.B., and Strachan, I. 1971. A correlation of Silurian rocks in the British Isles. Journal of the Geological Society of London, 127:103136.CrossRefGoogle Scholar
Copper, P. 1977. Paleolatitudes in the Devonian of Brazil and the Frasnian-Famennian mass extinctions, Palaeogeography, Palaeoclimatology, Palaeoecology, 21:165207.CrossRefGoogle Scholar
Copper, P. 1986. Frasnian-Famennian mass extinction and cold water oceans. Geology, 14:835839.2.0.CO;2>CrossRefGoogle Scholar
Copper, P. 1994a. Ancient reef ecosystem expansion and collapse. Coral Reefs, 13:311.CrossRefGoogle Scholar
Copper, P. 1994b. Reefs under stress: the fossil record. In Oekentorp, K. (ed.), Proceedings of the International Fossil Coral and Reef Symposium, Munster. Courier Forschungsinstitut Senckenberg, 172:8794.Google Scholar
Copper, P. 1996. Davidsonia and Rugodavidsonia (new genus), cryptic Devonian atrypid brachiopods from Europe and South China . Journal of Paleontology, 70:588602.CrossRefGoogle Scholar
Copper, P. 1997. Reefs and carbonate productivity: Cambrian through Devonian. Proceedings 8th International Coral Reef Symposium, Panama, 2:16231630.Google Scholar
Copper, P. 2001a. Reefs during the multiple crises towards the Ordovician - Silurian boundary: Anticosti Island, eastern Canada, and worldwide. Canadian Journal Earth Sciences, 38:153171.CrossRefGoogle Scholar
Copper, P. 2001b. Evolution, radiations, and extinctions in Proterozoic to Mid-Paleozoic reefs. In Stanley, G.D. Jr. (ed.), The history and sedimentology of ancient reef systems. Topics in Geobiology, 17:89119.CrossRefGoogle Scholar
Copper, P. 2002a. Reef development at the Frasnian-Famennian (Late Devonian) mass extinction boundary. In Racki, G., and House, M.R. (eds.), Late Devonian biotic crisis: ecological, depositional and geochemical records. Special Issue Palaeogeography, Palaeoclimatology, Palaeoecology, 181:2766.CrossRefGoogle Scholar
Copper, P. 2002b. Silurian and Devonian reefs: 80 million years of global greenhouse between two ice ages. In Kiessling, W., Flügel, E., and Golonka, J. (eds.), Phanerozoic reef patterns. SEPM Special Publications, 72:181238.Google Scholar
Copper, P., and Edinger, E. 2009. Distribution, geometry and palaeogeography of the Frasnian (Late Devonian) reef complexes of Banks Island, NWT, western arctic, Canada. In Königshof, P. (ed.) Devonian change: case studies in palaeogeography and palaeoecology. The Geological Society, London, Special Publications, 314:107122.CrossRefGoogle Scholar
Copper, P., and Scotese, C.R. 2003. Megareefs in Middle Devonian supergreenhouse climates. In Chan, M.A., and Archer, A.W. (eds.), Extreme depositional environments: mega end members in geologic time. Geological Society of America Special Paper, 370:209230.Google Scholar
Darwin, C. 1851. Geological observations on coral reefs, volcanic islands and on South America. Smith, Elder & Co., London, 250 pp.Google Scholar
Desrochers, A., Farley, C., Achab, A., Asselin, E., and Riva, J.F. 2010. A far-field record of the end Ordovician glaciation: the Ellis Bay Formation, Anticosti island, eastern Canada. Palaeogeography, Palaeoclimatology, Palaeoecology, 296:248263.CrossRefGoogle Scholar
Edinger, E.N., Copper, P., Risk, M.J., and Atmojo, W. 2002. Oceanography and reefs of Recent and Paleozoic tropical epeiric seas. Facies, 47:127150.CrossRefGoogle Scholar
Finnegan, S., Bergmann, K., Eiler, J.M., Jones, D.S., Fike, D.A., Eisenman, I., Hughes, N.C., Tripati, A.K., and Fischer, W.W. 2011. The magnitude and duration of Late Ordovician-Early Silurian glaciation. Science, 331:903906.CrossRefGoogle ScholarPubMed
Flügel, E. 2002. Triassic reef patterns. In Kiessling, W., Flügel, E., and Golonka, J. (eds.), Phanerozoic reef patterns. SEPM Special Publications, 72:391463.Google Scholar
Flügel, E., and Kiessling, W. 2002. Patterns of Phanerozoic reef cries. In Kiessling, W., Flügel, E., and Golonka, J. (eds.), Phanerozoic reef patterns. SEPM Special Publications, 72:690733.Google Scholar
Geldsetzer, H.H.J., James, N.P., and Tebbutt, G.E. (eds.). 1988. Reefs Canada and adjacent areas. Memoirs Canadian Society of Petroleum Geologists, 13:1775.Google Scholar
Ghienne, J.F. 2003. Late Ordovician sedimentary environments, glacial cycles and post-glacial transgression in the Taoudeni Basin, West Africa. Palaeogeography, Palaeoclimatology, Palaeoecology 189:117145.CrossRefGoogle Scholar
Ghienne, J.F., Boumendiel, K., Paris, F., Videt, B., Racheboeuf, P., and Salem, H.A. 2007. The Cambro-Ordovician succession in the Ougarta Range (western Algeria, North Africa) and interference of the Late Ordovician glaciation on the development of the Lower Palaeozoic transgression on northern Gondwana. Bulletin of Geosciences, 82:183214.CrossRefGoogle Scholar
Gradstein, F., Ogg, J., and Smith, A. (eds.). 2005. A geologic time scale 2004, Cambridge University Press, Cambridge, 589 pp.CrossRefGoogle Scholar
Gradstein, F., 2008. International Geologic Timescale. Newsletters on Stratigraphy, 43:p.9.Google Scholar
Herbert, T.D., Peterson, L.C., Lawrence, K.T., and Liu, Z.H. 2010. Tropical ocean temperatures over the past 3.5 million years. Science, 328:15301534.CrossRefGoogle ScholarPubMed
Holland, C.H., and Bassett, M.G. (eds.). 2002. Telychian rocks of the British Isles and China (Silurian, Llandovery Series), National Museums and Galleries of Wales, Geological Series, 21:1210.Google Scholar
Hoorn, C. and 16 co-authors. 2010. Amazonia through time: Andean uplift, climate change, landscape evolution, and biodiversity. Science, 330:927931.CrossRefGoogle ScholarPubMed
House, M.R. 2002. Strength, timing, setting and cause of mid-Palaeozoic extinctions. In Racki, G., and House, M.R. (eds.), Late Devonian biotic crisis: ecological, depositional and geochemical records. Special Issue Palaeogeography, Palaeoclimatology, Palaeoecology, 181:525.CrossRefGoogle Scholar
Hubmann, B., and Suttner, T. 2007. Siluro-Devonian Alpine reefs and pavements. In Alvaro, J.J., Aretz, M., Boulvain, F., Munnecke, A., Vachard, D., and Vennin, E. (eds.), Palaeozoic reefs and bioaccumulations: climatic and evolutionary controls. Geological Society, London Special Publications, 275:95108.CrossRefGoogle Scholar
Hubmann, B., Suttner, T., and Messner, F. 2006. Geologic framework of Palaeozoic reefs in Austria with special emphasis on Devonian reef-architecture of the Graz Palaeozoic. Joannea Geologie und Paläontologie, 8:4772.Google Scholar
Joachimski, M. M., Breisig, M.S., Buggisch, W., Talent, J.A., Mawson, R., Gereke, M., Morrow, J.R., Day, J., and Weddige, K. 2009. Devonian climate and reef evolution: insights from oxygen isotopes in apatite. Earth and Planetary Letters, 284:599609.CrossRefGoogle Scholar
Kaljo, D., Hints, L., Männick, P., and Nolvak, J. 2008. The succession of Hirnantian events based on data from Baltica: brachiopods, chitinozoans, conodonts and carbon isotopes. Estonian Journal of Earth Sciences, 57:197218.CrossRefGoogle Scholar
Kiessling, W, Aberhan, M., and Villier, L. 2008. Phanerozoic trends in skeletal mineralogy. Nature Geoscience, 1:527530.CrossRefGoogle Scholar
Kiessling, W., Flügel, E., and Golonka, J. 2000. Fluctuations in the carbonate production of Phanerozoic reefs. Geological Society, London, Special Publications, 178:191215.CrossRefGoogle Scholar
Kiessling, W., Flügel, E. and Golonka, J. 2002. Phanerozoic reef patterns. SEPM Special Publication, 72:1775.Google Scholar
Kiessling, W., Flügel, E., and Golonka, J. 2003. Patterns of Phanerozoic carbonate platform sedimentation. Lethaia, 36:195226.CrossRefGoogle Scholar
Kiessling, W., Simpson, C., and Foote, M. 2010. Reefs as cradles of evolution and sources of biodiversity in the Phanerozoic. Science, 327:196198.CrossRefGoogle ScholarPubMed
Kleypas, J.A., Buddemeier, R.W., Archer, D., Gattuso, J.P., Langdon, C., and Opdyke, B.N. 1999. Geochemical consequences of increased carbon dioxide on coral reefs. Science, 284:118120.CrossRefGoogle ScholarPubMed
Königshof, P. (ed.). 2009. Devonian change: case studies in palaeogeography and palaeoecology. Geological Society, London, Special Publications, 314:1287.CrossRefGoogle Scholar
Lacis, A.A., Schmidt, G.A., Rind, D., and Ruedy, R.A. 2010. Atmospheric CO2: principal control knob governing Earth's temperature. Science, 330:356359.CrossRefGoogle ScholarPubMed
Landing, E. and Johnson, M.E. (eds.). 2003. Silurian lands and seas- paleogeography outside of Laurentia. New York State Museum Bulletin, 493:1400.Google Scholar
Linsley, B.K. 1996. Oxygen isotope record of sealevel and climate variations in the Sulu Sea over the past 150,000 years. Nature, 380:234237.CrossRefGoogle Scholar
Long, D.G.F., and Copper, P. 1987. Stratigraphy of the Upper Ordovician upper Vaureal and Ellis Bay formations, eastern Anticosti island, Quebec. Canadian Journal of Earth Sciences, 24:18071820.CrossRefGoogle Scholar
McLean, R.A. 2010. Frasnian (Upper Devonian) colonial disphyllid corals from western Canada: taxonomy and biostratigraphic significance. NRC Research Press, Ottawa, Ont., 189 pp.Google Scholar
Mistiaen, B. 1984. Disparition des stromatopores paléozoiques ou survie du groupe: hypothèse et discussion. Bulletin Société Géologique de la France, 24:12451250.CrossRefGoogle Scholar
Mistiaen, B. 1994. Skeletal density: implication for development and extinction of Paleozoic stromatoporoids. Courier Forschungs-Institut Senckenberg, 172:319327.Google Scholar
Mistiaen, B. 2005. A variation of the skeletal density in Late Palaeozoic stromatoporoids: a response to climatic changes. Abstracts, ‘Climatic and evolutionary controls on Palaeozoic reefs and bioaccumulations’, Museum National d'Histoire Naturelle, Paris, 7–9 September 2005, p. 41.Google Scholar
Mistiaen, B., Milhau, B., Khatir, A., Hongfei, H., Vachard, D., and Xiantao, W. 1998. Famennien terminal (Strunien) d'Etroeungt (Avesnois, nord de la France) et d'Etaoucun (Guangxi, Chine du sud). Incidences paléogéographiques des données relatives aux stromatopores et ostracodes Annales de la Société Géologique du Nord, 6:97104.Google Scholar
Nestor, H., Copper, P., and Stock, C.W. 2010. Late Ordovician and Early Silurian stromatoporoid sponges from Anticosti Island, eastern Canada: crossing the O/S mass extinction boundary. NRC Research Press, Ottawa, Canada, 163 pp.Google Scholar
Puceat, F., Joachimski, M.M., Bouilloux, A., Monna, F., Bonion, A., MotreuiL, S., Morinière, F., Hénard, S., Mourin, J., Dera, G., and Quesne, D. 2010. Revised phosphate-water fractionation equation reassessing paleotemperatures derived from biogenic apatite. Earth and Planetary Science Letters, 298:135142.CrossRefGoogle Scholar
Rull, V. 2011. Origins of biodiversity, Science, 331:398399.CrossRefGoogle ScholarPubMed
Schuchert, C., and Twenhofel, W.H. 1910. Ordovicic-Siluric section of the Mingan and Anticosti islands, Gulf of St Lawrence. Bulletin Geological Society of America, 21:677716.CrossRefGoogle Scholar
Sepkoski, J. J. 1996. Patterns of Phanerozoic extinction: a perspective from global databases, pp. 3551 In Walliser, O. H. (ed.). Global events and event stratigraphy in the Phanerozoic. Springer Verlag, Berlin.CrossRefGoogle Scholar
Sepkoski, J. J. 2002. A compendium of fossil marine animal genera. Bulletin of American Paleontology, 363:1560.Google Scholar
Sorauf, J.E., and Peddler, A.E.H. 1986. Late Devonian rugose corals and the Frasnian-Famennian crisis. Canadian Journal of Earth Sciences, 23:12651287.CrossRefGoogle Scholar
Spalding, M.D., Ravilious, C., and Green, E. 2001. World atlas of coral reefs, University of California Press, Berkeley, USA, 424 pp.Google Scholar
Stanley, S. M. 1988. Phanerozoic mass extinctions: shared patterns suggest global cooling as a common cause. American Journal of Science, 288:344352.CrossRefGoogle Scholar
Stanley, S. M. 2006. Influence of seawater chemistry on biomineralization throughout Phanerozoic time: paleontological and experimental evidence. 232:214236.Google Scholar
Stearn, C.W. 1987. The effect of the Frasnian-Famennian extinction event on the stromatoporoids. Geology, 15:677680.2.0.CO;2>CrossRefGoogle Scholar
Tomasczik, T., Van Woese, R., and Mah, A.J. 1996. Rapid coral colonization of a lava flow following volcanic eruption, Banda Islands. Coral Reefs, 15:169175.CrossRefGoogle Scholar
Veron, J.E.N. 2008. Mass extinctions and ocean acidification: biological constraints or geological dilemmas. Coral Reefs, 27:459472.CrossRefGoogle Scholar
Veron, J.E.N., Hoegh-Guildberg, O., Lenton, T.M., Lough, J.M., Obura, D.O., Pearce-Kelly, P., Sheppard, C.R.C., Spalding, M., Stafford-Smith, M.G., and Rogers, A.D. 2009. The coral reef crisis: the critical importance of <350 ppmCO2. Marine Pollution Bulletin, 58:14281436.CrossRefGoogle Scholar
Wallace, A.R. 1880. Island life: or the phenomena and cause of insular faunas and floras including a revision and attempted solution of the problem of geological climates, MacMillan & Co., London, 526 pp.Google Scholar
Waters, J.A. and Webster, G.D. 2009. A re-evaluation of Famennian echinoderm diversity: implications for patterns of extinction and rebound in the Late Devonian. In Königshof, P. (ed.), Devonian change: case studies in palaeogeography and palaeoecology. Geological Society, London, Special Publications, 314:149161.CrossRefGoogle Scholar
Webby, B.D. 2002 In Kiessling, W., Flügel, E., and Golonka, J. (eds.), Phanerozoic reef patterns. SEPM Special Publications, 72:129179.Google Scholar
Webby, B.D. 2004. Stromatoporoids, pp. 112118 In Webby, B. D., Paris, F., Droser, M.L., and Percival, I.G. (eds.), The great Ordovician biodiversification event, Columbia University Press, New York, 484 pp.CrossRefGoogle Scholar
Webby, B.D., Elias, R.J., Young, G.A., Neuman, B.E.E., and Kaljo, D. 2004. Corals, pp. 124146 In Webby, B. D., Paris, F., Droser, M.L., and Percival, I.G. (eds.), The great Ordovician biodiversification event, Columbia University Press, New York, 484 pp.CrossRefGoogle Scholar
Wilder, R. 1989. Neue Ergebnisse zum oberdevonischen Riffsterben am Nordrand des mitteleuropäischen Variscikums. Fortschritte der Geologic von Rheinland und Westfalen, 35:5774.Google Scholar
Wilder, R. 1994. Death of Devonian reefs - implications and further investigations. Courier Forschungsinstitut Senckenberg, 172:241247.Google Scholar
Williams, A. 1972. Introduction and general aspects of correlation. In Williams, A., Strachan, A. I., Bassett, D.A., Dean, W.T., Ingham, J.K., Wright, A.D., and Whittington, H.B. (eds.), A correlation of Ordovician rocks of the British Isles, Geological Society of London, Special Report, 3:110.Google Scholar
Young, G.A., and Kershaw, S. 2005. Classification and controls of internal banding in Palaeozoic stromatoporoids and colonial corals. Palaeontology, 48:623651.CrossRefGoogle Scholar
Zeng, D., Bingwen, L., and Yunming, H. 1992. Reefs through geological ages in China, China Petroleum Industry Press, Beijing, 104, pp.Google Scholar
Ziegler, A.M. 1965. Silurian marine communities and their environmental significance. Nature, 207:270272.CrossRefGoogle Scholar